Investigation of Group 12 Metal Complexes with a Tridentate SNS

Investigation of Group 12 Metal Complexes with a Tridentate SNS Ligand by X-ray ... Department of Chemistry, The College of William and Mary, Williams...
0 downloads 0 Views 223KB Size
Inorg. Chem. 2006, 45, 571−581

Investigation of Group 12 Metal Complexes with a Tridentate SNS Ligand by X-ray Crystallography and 1H NMR Spectroscopy Wei Lai, Steven M. Berry, and Deborah C. Bebout* Department of Chemistry, The College of William and Mary, Williamsburg, Virginia 23187-8795 Raymond J. Butcher Department of Chemistry, Howard UniVersity, Washington, D.C. 20059 Received June 30, 2005

Two series of zinc triad complexes containing the ligand 2,6-bis(methylthiomethyl)pyridine (L1) were synthesized and characterized by X-ray crystallography and solution-state 1H NMR spectroscopy. The distorted meridional octahedral M(L1)2(ClO4)2 series includes the first structurally characterized Zn(II) and Cd(II) complexes with N2(SR2)4 coordination spheres. Coordination of HgCl2 and ZnCl2 with 1 equiv of ligand afforded mononuclear, five-coordinate species Hg(L1)Cl2 and Zn(L1)Cl2, respectively, with distorted square-pyramidal and trigonal-bipyramidal geometries. With CdCl2, the dimeric [Cd(L1)Cl(µ-Cl)]2 complex was obtained. The distorted octahedral coordination geometry of each Cd(II) center in this complex is formed by one tridentate ligand, two bridging chloride ions, and one terminal chloride ion. NMR spectra indicate that the intermolecular ligand-exchange rate of [M(L1)2]2+ decreased in the order Cd(II) > Zn(II) > Hg(II). Slow intermolecular ligand-exchange conditions on the chemical-shift time scale were found for 1:2 metal-to-ligand complexes of L1 with Hg(II) and Zn(II) but not Cd(II). Slow intermolecular ligandexchange conditions in acetonitrile-d3 solutions permitting detection of 3-5J(199Hg1H) were found for 1:1 and 1:2 Hg(ClO4)2/L1 complexes, but not for the related Cd(ClO4)2 complexes. The magnitudes of J(199Hg1H) for equivalent protons were smaller in [Hg(L1)2]2+ than in [Hg(L1)(NCCH3)x]2+. The relative intermolecular ligand-exchange rates of the zinc triad complexes investigated here suggest that the toxicity of Hg(II) is accentuated by the relative difficulty of displacing it from the coordination sites encountered.

Introduction The considerable differences between the bioactivities of the group 12 metal ions motivate parallel investigations of their coordination chemistry with biologically relevant ligands. In physiological systems, Zn(II) serves important catalytic and structural roles by selectively binding to sites of specific metalloproteins. On the other hand, Cd(II) and Hg(II) are toxic, and the molecular mechanisms for their toxicities are incompletely understood. The only beneficial role for Cd(II) is the recently discovered Cd(II)-specific carbonic anhydrase in the marine diatom Thalassiosira weissflogii, produced in the presence of Cd(II) under zinclimited conditions.1 Other than binding the transcription regulator of mercury detoxification proteins, no physiologi* Corresponding author. Phone: (757) 221-2558. Fax: (757) 221-2715. E-mail: [email protected]. (1) Lane, T. W.; Morel, F. M. M. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 4627.

10.1021/ic051091+ CCC: $33.50 Published on Web 12/20/2005

© 2006 American Chemical Society

cally beneficial roles have been discovered for Hg(II). The transcriptional regulation and metal binding of metalloproteins are highly selective for a single physiologically essential metal ion under normal circumstances. However, trace nonessential metal ions interfere with normal function. The potential toxicological targets of Cd(II) and Hg(II) are extensive because of their low coordination geometry and number preferences. Clarification of the range of slow intermolecular exchange binding environments for divalent zinc triad metal ions is required to fully appreciate the intricate coordination of specific metal-binding events in biochemistry and the most detrimental physiological effects of heavy metal ions. Investigating the coordination chemistry of the group 12 metal ions poses some unique challenges. The divalent forms of these metals are d10. Therefore, their coordination number and geometry preferences are weak compared to those of transition metal ions with incomplete outer d shells. As a Inorganic Chemistry, Vol. 45, No. 2, 2006

571

Lai et al.

Figure 1. Pyridine-based dithioether ligands.

result, their coordination chemistry with simple ligands is typically plagued by rapid intermolecular ligand exchange between a variety of complexes. Furthermore, the divalent group 12 metal ions are diamagnetic, and few spectroscopies are available for studying their coordination. Perhaps the most versatile spectroscopic method for characterizing organic complexes of these metals is NMR spectroscopy, which can detect slow intermolecular ligand exchange on the chemical-shift time scale for Zn(II), Cd(II), and Hg(II) complexes and on the J(111/113Cd1H) and J(199Hg1H) time scales for Cd(II) and Hg(II) complexes, respectively. Complexes of the group 12 metal ions with sulfur/nitrogen mixed donor ligands, such as thiolates or macrocyclic thioethers, have been widely investigated.2-6 Complexes with acyclic thioether ligands, however, have been less well studied. We recently structurally characterized seven coordination complexes of group 12 metals with the NSN-donor ligand bis(2-methylpyridyl) sulfide (Py2S).7 In the present article, X-ray crystallographic and solution NMR spectroscopic characterization of the coordination chemistry of divalent group 12 metal ions with the SNS-donor ligand 2,6bis(methylthiomethyl)pyridine (L1) is reported. As summarized in Figure 1, several other dithioether-derivatized lutidyl ligands have been reported.8-16 In crystallographically (2) Anjali, K. S.; Vittal, J. J.; Dean, P. A. W. Inorg. Chim. Acta 2003, 351, 79. (3) Bell, N. A.; Branston, T. N.; Clegg, W.; Creighton, J. R.; CucurullSanchez, L.; Elsegood, M. R. J.; Raper, E. S. Inorg. Chim. Acta 2000, 303, 220. (4) Castineiras, A.; Hiller, W.; Straehle, J.; Bravo, J.; Casas, J. S.; Gayoso, M.; Sordo, J. J. Chem. Soc., Dalton Trans. 1986, 1945. (5) Spencer, D. J. E.; Blake, A. J.; Parsons, S.; Schroder, M. J. Chem. Soc., Dalton Trans. 1999, 1041. (6) Brand, U.; Vahrenkamp, H. Inorg. Chem. 1995, 34, 3285. (7) Berry, S. M.; Bebout, D. C.; Butcher, R. J. Inorg. Chem. 2005, 44, 27. (8) Ball, R. J.; Genge, A. R. J.; Radford, A. L.; Skelton, B. W.; Tolhurst, V.-A.; White, A. H. J. Chem. Soc., Dalton Trans. 2001, 2807. (9) Marquez, V. E.; Anacona, J. R.; Hurtado, R. J.; De Delgado, G. D.; Roque, E. M. Polyhedron 1999, 18, 1903. (10) Canovese, L.; Visentin, F.; Chessa, G.; Uguagliati, P.; Santo, C.; Bandoli, G.; Maini, L. Organomet. 2003, 22, 3230. (11) Vinas, C.; Angles, P.; Sanchez, G.; Lucena, N.; Teixidor, F.; Escriche, L.; Casabo, J.; Piniella, J. F.; Alvarez-Larena, A.; Kivekaes, R.; Sillanpaeae, R. Inorg. Chem. 1998, 37, 701. (12) Canovese, L.; Visentin, F.; Uguagliati, P.; Lucchini, V.; Bandoli, G. Inorg. Chim. Acta 1998, 277, 247. (13) Teixidor, F.; Sanchez-Castello, G.; Lucena, N.; Escriche, L.; Kivekas, R.; Sundberg, M.; Casabo, J. Inorg. Chem. 1991, 30, 4931. (14) Teixidor, F.; Escriche, L.; Rodriguez, I.; Casabo, J.; Rius, J.; Molins, E.; Martinez, B.; Miravitlles, C. J. Chem. Soc., Dalton Trans. 1989, 1381. (15) Escriche, L.; Sanz, M.; Casabo, J.; Teixidor, F.; Molins, E.; Miravitlles, C. J. Chem. Soc., Dalton Trans. 1989, 1739.

572 Inorganic Chemistry, Vol. 45, No. 2, 2006

characterized metal complexes, these ligands generally behave as tridentate chelating ligands, giving two fivemembered rings.8,10,15,16 Bidentate binding modes have been observed in [Pd(Me)(L1)]12 and [Cu(L3)Br(µ-Br)]2,8 where only the pyridine nitrogen and one thioether moiety bind the metal ions. The only group 12 metal ion complexes previously reported with this family of ligands are [Zn(L2)Br2]16 and [Cd(L2)Cl(µ-Cl)]2.14 In this article, the synthesis and characterization of MCl2 and M(ClO4)2 [M ) Zn(II), Cd(II), Hg(II)] complexes of L1 are reported. X-ray crystallographic structures are presented for nonhomologous chloride complexes with 1:1 metal-toligand stoichiometry and for a homologous [M(L1)2](ClO4)2 series. These are the first complexes having the stoichiometry M(L)2 with this class of ligands. In addition, the Zn(II) and Cd(II) complexes with N2(SR2)4 coordination spheres are the first to be crystallographically characterized. Conditions for slowed intermolecular ligand exchange on the chemical-shift time scale in acetonitrile-d3 solution were found for [Zn(L1)2]2+ and [Hg(L1)2]2+, but not for [Cd(L1)2]2+. In addition, slow intermolecular exchange conditions on the J(M1H) time scale were found for [Hg(L1)2]2+ and [Hg(L1)(CH3CN)x]2+, but not for comparable Cd(II) complexes. The lowest intermolecular exchange rates for this SNS ligand were observed with Hg(II), followed by Zn(II) and finally Cd(II). This observation suggests that the relative difficulty of displacing Hg(II) from mixed nitrogen/thioether sulfur coordination environments contributes to its broad toxicological effects. Experimental Section Solvents and reagents were of commercial grade and were used as received unless otherwise stated. Elemental analyses were carried out by Atlantic Microlab, Inc., Norcross, GA. The proton labeling scheme is shown in Figure 1. Reported coupling constants are 3J(1H1H) unless otherwise noted. Caution! All of the perchlorate salts of metal complexes included in this work were stable for routine synthesis and purification procedures. HoweVer, organic perchlorates are potentially explosiVe and should be handled with care.17 Synthesis of 2,6-Bis(methylthiomethyl)pyridine (L1). The ligand was prepared by a variation of the procedure described by Canovese and co-workers.18 Sodium methanethiolate (3.1 g, 44 mmol) was added to an ice-cooled solution of 2,6-bis(bromomethyl)pyridine19 (5.3 g, 20 mmol) in DMF (50 mL). The mixture was brought to room temperature and stirred for 24 h. The solvent was removed by Kugelrohr distillation (50 °C, 0.5 mmHg). Continued Kugelrohr distillation (150 °C, 0.01 mmHg) of the yellow oily residue yielded L1 as a colorless oil (3.3 g, 83%). 1H NMR (CD3CN, 20 °C): δ 7.70 (t, 1H, J ) 8 Hz, Hp), 7.26 (d, 2H, J ) 8 Hz, Hm), 3.75 (s, 4H, Hf), 2.07(s, 6H, Me). Anal. Calcd for C9H13NS2: C, 27.09; H, 3.28; N, 3.51. Found: C, 27.09; H, 3.27; N, 3.53. Preparation of Hg(L1)2(ClO4)2 (1). A solution of Hg(ClO4)2‚ 3H2O (230 mg, 0.50 mmol) in 5 mL of methanol was added dropwise to a solution of L1 (100 mg, 0.5 mmol) in 5 mL of (16) Teixidor, F.; Escriche, L.; Casabo, J.; Molins, E.; Miravitlles, C. Inorg. Chem. 1986, 25, 4060. (17) Raymond, K. N. Chem. Eng. News 1983, 61, 4. (18) Canovese, L.; Chessa, G.; Marangoni, G.; Pitteri, B.; Uguagliati, P.; Visentin, F. Inorg. Chim. Acta 1991, 186, 79. (19) Baker, W.; Buggle, K. M.; McOmie, J. F. W.; Watkins, D. A. M. J. Chem. Soc. 1958, 3594.

Zn Triad Coordination of 2,6-Bis(methylthiomethyl)pyridine Table 1. Crystallographic Data for Complexes 1-3

empirical formula fw crystal system space group a, Å b, Å c, Å R, deg β, deg γ, deg V, Å3 Z dcalc, g cm-3 µ, mm-1 T, K R1,a wR2b [I > 2σ(I)] R1,a wR2b (all data) a

1

2

3

C18H26N2S4HgCl2O8 798.14 orthorhombic Ibca 13.5814(17) 14.184(3) 27.412(3) 90 90 90 5280.5(14) 8 2.008 6.394 148(2) 0.0452, 0.1026 0.0597, 0.1115

C18H26N2S4CdCl2O8 709.95 orthorhombic Pcca 14.2898(19) 13.5492(17) 27.536(4) 90 90 90 5331.4(12) 8 1.769 1.378 103(2) 0.0234, 0.0582 0.0377, 0.0613

C18H26N2S4ZnCl2O8 662.92 monoclinic P21/n 11.982(6) 11.667(6) 19.559(11) 90 103.672(9) 90 2657(2) 4 1.657 1.485 103(2) 0.0396, 0.0974 0.0605, 0.1080

R1 ) ∑||Fo| - |Fc||/∑|Fo|. b wR2 ) ∑[w(Fo2 - Fc2)2]/∑[w(Fo2)2]1/2.

methanol with stirring. A white precipitate formed. The solution was filtered, and the filtrate was diluted slowly with 10 mL of acetonitrile and set aside for slow evaporation. Colorless X-rayquality crystals of 1 (91 mg, 46%) formed upon standing for 4 days. Mp: 170-171 °C (dec). 1H NMR (2 mM, CD3CN, 20 °C): δ 8.15 [t, 1H, J ) 8 Hz, 5J(199Hg1H) ) 15 Hz, Hp], 7.54 [d, 2H, J ) 8 Hz, 4J(199Hg1H) ) 15 Hz, Hm], 4.15 [s, 4H, 3J(199Hg1H) ) 39, Hf], 2.18 [s, 6H, 3J(199Hg1H) ) 49 Hz, Me]. Anal. Calcd for C18H26Cl2HgN2O8S4: C, 54.26; H, 6.53; N, 7.03. Found: C, 54.43; H, 6.64; N, 7.01%. Preparation of Cd(L1)2(ClO4)2 (2). A solution of Cd(ClO4)2‚ 6H2O (210 mg, 0.50 mmol) in 5 mL of acetonitrile was added dropwise to a solution of L1 (100 mg, 0.5 mmol) in 5 mL of methanol with stirring. A white precipitate formed. The solution was clarified by the addition of 5 mL of acetonitrile. The mixture was slowly diluted with 6 mL of toluene and set aside for slow evaporation. Colorless X-ray-quality crystals of 2 (100 mg, 60%) formed upon standing for 4 days. Mp: 248-250 °C (dec). 1H NMR (2 mM, CD3CN, 20 °C): δ 8.01 (t, 1H, J ) 8 Hz, Hp), 7.54 (d, 2H, J ) 8 Hz, Hm), 4.15 (s, 4H, Hf), 2.13 (s, 6H, Me). Anal. Calcd for C18H26CdCl2N2O8S4: C, 30.46; H, 3.66; N, 3.95. Found: C, 30.41; H, 3.63; N, 3.88. Preparation of Zn(L1)2(ClO4)2 (3). A solution of Zn(ClO4)2‚ 3H2O (190 mg, 0.50 mmol) in 3 mL of acetone was added dropwise to a solution of L1 (100 mg, 0.5 mmol) in 2 mL of acetone with stirring. Colorless crystals of 3 (26 mg, 15%) suitable for X-ray crystallography were obtained in 5-mm glass tubes after 5 days by slow diffusion of the mixture into toluene. Mp: 229-231 °C. 1H NMR (2 mM, CD3CN, 20 °C): δ 7.99 (t, 1H, J ) 8 Hz, Hp), 7.51 (d, 2H, J ) 7 Hz, Hm), 4.07 (s, 4H, Hf), 2.10 (s, 6H, Me). Anal. Calcd for C18H26Cl2N2O8S4Zn: C, 32.61; H, 3.95; N, 4.23. Found: C, 31.86; H, 4.04; N, 4.18. Preparation of Hg(L1)Cl2 (4). A solution of HgCl2 (140 mg, 0.50 mmol) in 5 mL of methanol was added dropwise to a solution of L1 (100 mg, 0.5 mmol) in 5 mL of acetonitrile with stirring. Slow evaporation of the solution resulted in the formation of colorless X-ray-quality crystals of 4 (90 mg, 40%) in 10 days. Mp: 132-133 °C. 1H NMR (2 mM, CD3CN, 20 °C): δ 7.93 (t, 1H, J ) 8 Hz, Hp), 7.46 (d, 2H, J ) 8 Hz, Hm), 4.14 (s, 4H, Hf), 2.05 (s, 6H, Me). Anal. Calcd for C9H13Cl2HgNS2: C, 22.96; H, 2.78; N, 2.98. Found: C, 23.05; H, 2.75; N, 2.91. Preparation of [Cd(L1)Cl(µ-Cl)]2 (5). A solution of CdCl2 (90 mg, 0.5 mmol) in 10 mL of methanol was added dropwise to a solution of L1 (100 mg, 0.5 mmol) in 5 mL of methanol with

stirring. The precipitate was filtered off and dried in vacuo (130 mg, 67%). Recrystallization from acetonitrile and toluene gave colorless X-ray-quality crystals of 5. The complex decomposed at 205 °C. 1H NMR (CD3CN, 20 °C): δ 7.96 (t, 1H, J ) 8 Hz, Hp), 7.48 (d, 2H, J ) 8 Hz, Hm), 4.15 (s, 4H, Hf), 2.06 (s, 6H, Me). Anal. Calcd for C18H26Cd2Cl4N2S4: C, 28.25; H, 3.42; N, 3.66. Found: C, 28.32; H, 3.39; N, 3.61. Preparation of Zn(L1)Cl2 (6). A solution of ZnCl2 (68 mg, 0.50 mmol) in 5 mL of acetone was added dropwise to a solution of L1 (100 mg, 0.5 mmol) in 5 mL of acetone with stirring. The solution was slowly diluted with 10 mL of toluene and set aside for slow evaporation. Colorless X-ray-quality crystals of 6 (110 mg, 65%) formed upon standing for 3 days. Mp: 198-199 °C. 1H NMR (2 mM, CD3CN, 20 °C): δ 7.96 (t, 1H, J ) 7 Hz, Hp), 7.48 (d, 2H, J ) 7 Hz, Hm), 4.05 (s, 4H, Hf), 2.27 (s, 6H, Me). Anal. Calcd for C9H13Cl2NS2Zn: C, 32.21; H, 3.90; N, 4.17. Found: C, 32.30; H, 3.79; N, 4.19. Solution-State NMR Spectroscopy. NMR spectra were collected in 5-mm-o.d. NMR tubes on a Varian Mercury 400VX NMR spectrometer operating in the pulse Fourier transform mode. All solutions for low-temperature NMR measurements were prepared using calibrated autopipets by dilution of metal salt and L1 stock solutions with CD3CN. The metal ion concentration in the solutions used in the NMR analysis was 2 mM. The sample temperature was maintained by blowing chilled nitrogen over the NMR tube in the probe. Proton chemical shifts were measured relative to internal solvent but are reported relative to tetramethylsilane (TMS). X-ray Crystallography. Selected crystallographic data are reported in Tables 1 and 2, and the complete data are provided in CIF files as Supporting Information. Data were collected on a Bruker SMART 1K CCD four-circle diffractometer using graphitemonochromated Mo KR X-radiation (λ ) 0.71073 Å). During data collection, 50 frames were measured at both the beginning and end of the data collection to monitor intensity decay. The structures were solved by direct methods and refined on F2 by full-matrix least-squares using the SHELXTL97 program package.20 All nonhydrogen atoms were fixed as anisotropic, the hydrogen atomic positions were fixed relative to the bonded carbons, and the isotropic thermal parameters were fixed. (20) Sheldrick, G. M. SHELXL, version 6.14; University of Go¨ttingen: Go¨ttingen, Germany, 1997.

Inorganic Chemistry, Vol. 45, No. 2, 2006

573

Lai et al. Table 2. Crystallographic Data for Complexes 4-6

empirical formula fw crystal system space group a, Å b, Å c, Å R, deg β, deg γ, deg V, Å3 Z dcalc, g cm-3 µ, mm-1 T, K R1,a wR2b [I > 2σ(I)] R1,a wR2b (all data) a

4

5

6

C9H13NS2HgCl2 470.81 triclinic P1h 7.57(2) 7.84(2) 11.67(3) 87.92(5) 83.39(5) 81.87(5) 681(3) 2 2.297 11.975 148(2) 0.0820, 0.2255 0.0877, 0.2320

C18H26N2S4Cd2Cl4 765.25 monoclinic P21 7.8108(13) 14.487(3) 11.835(2) 90 103.095(3) 90 1304.3(4) 2 1.948 2.371 148(2) 0.0332, 0.0862 0.0340, 0.0869

C9H13NS2ZnCl2 335.59 orthorhombic P212121 15.090(6) 15.121(6) 11.663(5) 90 90 90 2661.2(19) 8 1.675 2.529 103(2) 0.0172, 0.0405 0.0187, 0.0410

Figure 4. Structure of the cation of Zn(L1)2(ClO4)2 (3). Thermal ellipsoids are shown at the 50% level. Anions are omitted for clarity.

R1 ) ∑||Fo| - |Fc||/∑|Fo|. b wR2 ) ∑[w(Fo2 - Fc2)2]/∑[w(Fo2)2]1/2.

Figure 2. Structure of the cation of Hg(L1)2(ClO4)2 (1). Thermal ellipsoids are shown at the 50% level. Anions are omitted for clarity.

Figure 3. Structure of one of the two similar cations in the unit cell of Cd(L1)2(ClO4)2 (2). Thermal ellipsoids are shown at the 50% level. Anions are omitted for clarity.

Results and Discussion L1 acts as a tridentate chelating ligand in the solid-state structures of complexes 1-6. Complexes 1-3 consist of the dications [Hg(L1)2]2+, [Cd(L1)2]2+, and [Zn(L1)2]2+, respectively (Figures 2-4), and well-separated perchlorate anions. Selected bond distances and angles for complexes 1-3 are given in Table 3. Complexes 1-3 are the first crystallographically characterized complexes of this class of ligands with 1:2 metal-to-ligand ratios. The structures of chloride complexes Hg(L1)Cl2 (4), [Cd(L1)Cl(µ-Cl)]2 (5), and Zn(L1)Cl2 (6) are shown in Figures 5-7, respectively. Similar coor-

574 Inorganic Chemistry, Vol. 45, No. 2, 2006

dination spheres have been observed in [Cu(L8)Cl2],13 [Cu(L2)Cl2],15 [Cu(L3)Br], [Ni(L3)Br2],8 [Cd(L2)Cl(µ-Cl)]2,14 and [Zn(L2)Br2].16 Selected bond distances and angles for complexes 4-6 are given in Table 4. Crystal Structure of Hg(L1)2(ClO4)2 (1). As shown in Figure 2, the [Hg(L1)2]2+ cation has a meridional octahedral structure with crystallographically imposed C2 symmetry and a N-Hg-N bond angle of 174.7(2)°. The L1 ligands are rotated by 73° relative to each other, with intraligand S-Hg-S and average N-Hg-S bond angles of 149.16(5)° and 74.6(7)°, respectively. The thioether methyl groups of both ligands are arranged trans with respect to the pyridyl planes. The 2.406(5) Å Hg-N distance is similar to the HgNpyridyl distances observed in other six-coordinate complexes such as [Hg(pyridine)6](CF3SO3)2 and Hg(BMPA)2(ClO4)2 {BMPA ) bis[(2-pyridyl)methyl]amine}.21,22 The Hg-S distances of 2.676(2) and 2.754(2) Å are close to those [2.654(1) and 2.671(1) Å] recently reported for Hg(Py2S)2(ClO4)27 and comparable to those observed in six-coordinate Hg(II) complexes of macrocyclic thioethers.23,24 Each of the five-membered chelate rings in 1 adopts an envelope conformation with the S atom 1.0 Å from the plane containing Hg, N, and two lutidyl C atoms. The HgO(perchlorate) distances (3.87-6.35 Å) are greater than the sum of the van der Waals radii for Hg and O [rvdw(Hg) ) 1.70-2.00 Å, rvdw(O) ) 1.54 Å].25 The only other structurally characterized Hg(II) complex with a N2(SR2)4 metal coordination sphere is the Hg(PF6)2 complex of 1,4,10,13-tetrathia-7,16-diazacyclo-octadecane (TDO).26 This complex crystallized in a distorted octahedral coordination geometry. The average Hg-S distance for the four sulfur donors in the equatorial plane of the TDO (21) Bebout, D. C.; DeLanoy, A. E.; Ehmann, D. E.; Kastner, M. E.; Parrish, D. A.; Butcher, R. J. Inorg. Chem. 1998, 37, 2952. (22) Aakesson, R.; Sandstroem, M.; Staalhandske, C.; Persson, I. Acta Chem. Scand. 1991, 45, 165. (23) Marcus, S. T.; Bernhardt, P. V.; Grondahl, L.; Gahan, L. R. Polyhedron 1999, 18, 3451. (24) Blake, A. J.; Pasteur, E. C.; Reid, G.; Schroder, M. Polyhedron 1991, 10, 1545. (25) Nyburg, S. C.; Faerman, C. H. Acta Crystallogr. B: Struct. Sci 1985, B41, 274. (26) Blake, A. J.; Reid, G.; Schroeder, M. Polyhedron 1990, 9, 2931.

Zn Triad Coordination of 2,6-Bis(methylthiomethyl)pyridine Table 3. Selected Bond Distances (Å) and Bond Angles (deg) for Complexes 1-3 Hg(L1)2(ClO4)2 (1)a

Cd(L1)2(ClO4)2 (2)b

Zn(L1)2(ClO4)2 (3)

Hg-N Hg-S(1) Hg-S(2)

2.406(5) 2.6763(16) 2.7538(17)

Cd(1)-N(1A) Cd(1)-S(1A) Cd(1)-S(2A) Cd(2)-N(1B) Cd(2)-S(1B) Cd(2)-S(2B)

2.3558(14) 2.7077(5) 2.7356(6) 2.3673(15) 2.6819(5) 2.7359(5)

Zn-N(1A) Zn-S(1A) Zn-S(2A) Zn-N(1B) Zn-S(1B) Zn-S(2B)

2.125(3) 2.5730(14) 2.5519(13) 2.119(3) 2.5300(11) 2.6508(14)

N-Hg-S(1) N-Hg-S(2) N-Hg-N#1 N-Hg-S(1)#1 N-Hg-S(2)#1 S(1)-Hg-S(2) S(1)-Hg-S(1)#1 S(1)-Hg-S(2)#1 S(2)-Hg-S(2)#1

75.28(13) 73.88(13) 174.7(2) 107.95(12) 102.85(12) 149.16(5) 107.87(7) 81.20(5) 106.37(8)

N(1A)-Cd(1)-S(1A) N(1A)-Cd(1)-S(2A) N(1A)-Cd(1)-N(1A)#1 N(1A)-Cd(1)-S(1A)#1 N(1A)-Cd(1)-S(2A)#1 S(1A)-Cd(1)-S(2A) S(1A)-Cd(1)-S(1A)#1 S(1A)-Cd(1)-S(2A)#1 S(2A)-Cd(1)-S(2A)#1 N(1B)-Cd(2)-S(1B) N(1B)-Cd(2)-S(2B)#2 N(1B)-Cd(2)-N(1B)#2 N(1B)-Cd(2)-S(1B)#2 N(1B)-Cd(2)-S(2B) S(1B)-Cd(2)-S(2B) S(1B)-Cd(2)-S(1B)#2 S(1B)-Cd(2)-S(2B)#2 S(2B)-Cd(2)-S(2B)#2

75.69(4) 75.41(4) 178.02(7) 105.57(4) 103.36(4) 151.077(14) 104.51(2) 82.353(17) 105.37(2) 76.10(4) 75.15(4) 175.05(7) 106.99(4) 101.84(4) 151.168(14) 106.28(2) 80.299(17) 107.72(2)

N(1A)-Zn-S(1A) N(1A)-Zn-S(2A) N(1A)-Zn-N(1B) N(1A)-Zn-S(1B) N(1A)-Zn-S(2B) S(1A)-Zn-S(2A) S(1A)-Zn-N(1B) S(1A)-Zn-S(1B) S(1A)-Zn-S(2B) N(1B)-Zn-S(2A) S(1B)-Zn-S(2A) S(2A)-Zn-S(2B) N(1B)-Zn-S(1B) N(1B)-Zn-S(2B) S(1B)-Zn-S(2B)

79.56(7) 80.06(8) 175.23(9) 102.76(8) 97.20(8) 159.61(3) 98.10(7) 92.15(4) 88.66(4) 102.19(8) 92.83(4) 93.39(4) 81.41(8) 78.53(8) 159.84(3)

a Symmetry transformations used to generate equivalent atoms: -x, -y + 1/ , z; x, -y + 1, -z + 1/ ; -x - 1/ , y, -z. b Symmetry transformations used 2 2 2 to generate equivalent atoms: -x + 3/2, -y + 1, z; -x + 1/ 2, -y, z; -x + 1, y, -z + 3/2; -x + 2, y, -z + 3/2.

Figure 5. Structure of Hg(L1)Cl2 (4). Thermal ellipsoids are shown at the 50% level.

Figure 6. Structure of [Cd(L1)Cl(µ-Cl)]2 (5). Thermal ellipsoids are shown at the 50% level.

complex was 2.695(5) Å, which is very similar to the equatorial plane distance of complex 1. The aliphatic nitrogens in the axial positions had a Hg-N distance of

Figure 7. Structure of one of the two similar complexes in the unit cell of Zn(L1)Cl2 (6). Thermal ellipsoids are shown at the 50% level.

2.472(2) Å, slightly longer than the Hg-Npyridyl distance in 1. Crystal Structure of Cd(L1)2(ClO4)2 (2). Complex 2 crystallizes in the orthorhombic Pcca space group with eight formula units per cell. The unit cell contains two similar but crystallographically independent [Cd(L1)2]2+ cations (Figure 2). Both meridional octahedral [Cd(L1)]2+ cations have crystallographically imposed C2 symmetry. The N-M-N bond angles of 175.05(7)° and 178.02(7)° in 2 are closer to linear than are those in 1. The L1 pyridyl planes are rotated by 77° relative to each other. The average M-N [2.362(7) Å] and M-S [2.7153(3) Å] bond lengths of 2 are slightly shorter and nearly identical, respectively, to those in 1. The thioether methyl groups of both ligands are arranged trans with respect to the pyridyl planes, and the average intraligand S-M-S and N-M-S bond angles of 151.12(6)° and 75.591(5)° in 2 are similar to those observed Inorganic Chemistry, Vol. 45, No. 2, 2006

575

Lai et al. Table 4. Selected Bond Distances (Å) and Bond Angles (deg) for Complexes 4-6 Hg(L1)Cl2 (4)

[Cd(L1)Cl(µ-Cl)]2 (5)

Zn(L1)Cl2 (6)

Hg-Cl(1) Hg-Cl(2) Hg-N Hg-S(1) Hg-S(2)

2.383(7) 2.409(6) 2.454(11) 2.777(7) 2.854(7)

Cd(1)-N(1A) Cd(1)-Cl(2A) Cd(1)-Cl(1A) Cd(1)-Cl(2B) Cd(1)-S(1A) Cd(1)-S(2A) Cd(2)-N(1B) Cd(2)-Cl(1B) Cd(2)-Cl(2B) Cd(2)-S(1B) Cd(2)-S(2B) Cd(2)-Cl(2A)

2.395(4) 2.5301(12) 2.5355(11) 2.7050(11) 2.7060(12) 2.7342(12) 2.356(4) 2.5192(13) 2.5279(11) 2.7080(13) 2.7359(12) 2.7477(12)

Zn(1)-N(1A) Zn(1)-Cl(2A) Zn(1)-Cl(1A) Zn(1)-S(1A) Zn(1)-S(2A) Zn(2)-N(1B) Zn(2)-Cl(1B) Zn(2)-Cl(2B) Zn(2)-S(1B) Zn(2)-S(2B)

2.0785(15) 2.2333(8) 2.2548(7) 2.5741(9) 2.5814(9) 2.0828(15) 2.2190(8) 2.2326(8) 2.6168(9) 2.6318(8)

Cl(1)-Hg-Cl(2) Cl(1)-Hg-N Cl(2)-Hg-N Cl(1)-Hg-S(1) Cl(2)-Hg-S(1) N-Hg-S(1) Cl(1)-Hg-S(2) Cl(2)-Hg-S(2) N-Hg-S(2) S(1)-Hg-S(2)

139.31(17) 116.7(3) 103.9(3) 101.97(13) 93.20(16) 71.3(3) 94.89(15) 95.27(18) 71.4(3) 142.70(10)

N(1A)-Cd(1)-Cl(2A) N(1A)-Cd(1)-Cl(1A) Cl(2A)-Cd(1)-Cl(1A) N(1A)-Cd(1)-Cl(2B) Cl(2A)-Cd(1)-Cl(2B) Cl(1A)-Cd(1)-Cl(2B) N(1A)-Cd(1)-S(1A) Cl(2A)-Cd(1)-S(1A) Cl(1A)-Cd(1)-S(1A) Cl(2B)-Cd(1)-S(1A) N(1A)-Cd(1)-S(2A) Cl(2A)-Cd(1)-S(2A) Cl(1A)-Cd(1)-S(2A) Cl(2B)-Cd(1)-S(2A) S(1A)-Cd(1)-S(2A) N(1B)-Cd(2)-Cl(1B) N(1B)-Cd(2)-Cl(2B) Cl(1B)-Cd(2)-Cl(2B) N(1B)-Cd(2)-S(1B) Cl(1B)-Cd(2)-S(1B) Cl(2B)-Cd(2)-S(1B) N(1B)-Cd(2)-S(2B) Cl(1B)-Cd(2)-S(2B) Cl(2B)-Cd(2)-S(2B) S(1B)-Cd(2)-S(2B) N(1B)-Cd(2)-Cl(2A) Cl(1B)-Cd(2)-Cl(2A) Cl(2B)-Cd(2)-Cl(2A) S(1B)-Cd(2)-Cl(2A) S(2B)-Cd(2)-Cl(2A) Cd(1)-Cl(2A)-Cd(2) Cd(2)-Cl(2B)-Cd(1)

168.48(9) 93.51(10) 97.43(4) 84.45(10) 84.48(4) 177.01(4) 73.09(10) 109.93(4) 92.26(4) 89.23(4) 73.79(9) 103.11(4) 86.93(4) 90.40(4) 146.76(4) 91.26(10) 169.95(9) 97.33(4) 76.62(10) 101.50(4) 106.56(4) 74.00(9) 88.98(4) 100.84(4) 148.98(4) 87.66(10) 178.38(4) 83.64(4) 79.43(4) 89.56(4) 95.02(4) 96.13(4)

N(1A)-Zn(1)-Cl(2A) N(1A)-Zn(1)-Cl(1A) Cl(2A)-Zn(1)-Cl(1A) N(1A)-Zn(1)-S(1A) Cl(2A)-Zn(1)-S(1A) Cl(1A)-Zn(1)-S(1A) N(1A)-Zn(1)-S(2A) Cl(2A)-Zn(1)-S(2A) Cl(1A)-Zn(1)-S(2A) S(1A)-Zn(1)-S(2A) N(1B)-Zn(2)-Cl(1B) N(1B)-Zn(2)-Cl(2B) Cl(1B)-Zn(2)-Cl(2B) N(1B)-Zn(2)-S(1B) Cl(1B)-Zn(2)-S(1B) Cl(2B)-Zn(2)-S(1B) N(1B)-Zn(2)-S(2B) Cl(1B)-Zn(2)-S(2B) Cl(2B)-Zn(2)-S(2B) S(1B)-Zn(2)-S(2B)

117.21(4) 113.15(4) 129.62(2) 81.78(4) 88.89(3) 97.09(3) 79.79(4) 95.82(2) 93.96(3) 161.101(16) 119.52(5) 117.61(4) 122.87(2) 79.12(5) 98.68(2) 92.08(3) 79.96(5) 91.63(3) 97.55(3) 159.083(18)

in complex 1, indicating that the ligands are in a similar conformation. No other structurally characterized Cd(II) complex with a N2(SR2)4 metal coordination sphere has been reported. Crystal Structure of Zn(L1)2(ClO4)2 (3). As shown in Figure 4, the [Zn(L1)2]2+ cation in complex 3 has a more ideally meridional octahedral structure than the cations of complexes 1 and 2. Unlike for 1 and 2, there is no crystallographically imposed symmetry for 3. The 175.23(9)° N-M-N bond angle in 3 is comparable to those observed in 1 and 2. The L1 pyridyl planes are rotated by 87° relative to each other. The two methyl groups in the thioether moieties are placed in a trans fashion with respect to the pyridyl rings, as observed in complexes 1 and 2. The ZnNpyridyl distances of 2.119(3) and 2.125(3) Å are similar to those observed in six-coordinate Zn(II) complexes6,27-30 and (27) Parra-Hake, M.; Larter, M. L.; Gantzel, P.; Aguirre, G.; Ortega, F.; Somanathan, R.; Walsh, P. J. Inorg. Chem. 2000, 39, 5400. (28) Neels, A.; Stoeckli-Evans, H. Inorg. Chem. 1999, 38, 6164. (29) Glerup, J.; Goodson, P. A.; Hodgson, D. J.; Michelsen, K.; Nielsen, K. M.; Weihe, H. Inorg. Chem. 1992, 31, 4611. (30) Matsuda, K.; Takayama, K.; Irie, M. Inorg. Chem. 2004, 43, 482.

576 Inorganic Chemistry, Vol. 45, No. 2, 2006

also compare well with the Zn-N(His) distance in proteins,31-35 but they are significantly smaller than the M-N bond lengths in 1 and 2. The Zn-S bond distances ranging from 2.530(1) to 2.651(1) Å are comparable to those in sixcoordinate Zn(II) complexes,2,36 but also significantly shorter than all of the M-S bond lengths in 1 and 2. The average intraligand S-M-S and N-M-S bond angles of 159.72(8) and 79.89(8)°, respectively, in complex 3 are somewhat larger than those in complex 1 and 2 because of the shorter Zn-N bond lengths. No other structurally characterized Zn(II) complex with a N2(SR2)4 metal coordination sphere has been reported. (31) Becker, A.; Schlichting, I.; Kabsch, W.; Groche, D.; Schultz, S.; Wagner, A. F. V. Nat. Struct. Biol. 1998, 5, 1053. (32) Stec, B.; Hehir, M. J.; Brennan, C.; Nolte, M.; Kantrowitz, E. R. J. Mol. Biol. 1998, 277, 647. (33) Korkhin, Y.; Kalb, A. J.; Peretz, M.; Bogin, O.; Burstein, Y.; Frolow, F. J. Mol. Biol. 1998, 278, 967. (34) Fabiane, S. M.; Sohi, M. K.; Wan, T.; Payne, D. J.; Bateson, J. H.; Mitchell, T.; Sutton, B. J. Biochemistry 1998, 37, 12404. (35) Hyvonen, M.; Saraste, M. EMBO J. 1997, 16, 3396. (36) Helm, M. L.; Combs, C. M.; VanDerveer, D. G.; Grant, G. J. Inorg. Chim. Acta 2002, 338, 182.

Zn Triad Coordination of 2,6-Bis(methylthiomethyl)pyridine Table 5. Comparison of M-Naliphatic, M-Naromatic, and M-S Bond Length Ranges for Known Complexes with N2(SR2)4 Coordination Spheres metal ion Fe(II) Co(II) Co(III) Ni(II) Cu(II) Zn(II) Cd(II) Hg(II)

M-Naliphatic (Å)

M-Naromatic (Å)

2.022(4)-2.038(4) 2.100(3) 1.9993(4), 1.994(4) 2.065(13)-2.126(13) 2.007(13)-2.191(17) 2.472(17), 2.473(11)

2.053(2)-2.091(5) 2.119(3)-2.125(3) 2.3558(14)-2.3673(15) 2.406(5)

Comparisons of Complexes 1-3 to Related Complexes of Physiologically Essential Metal Ions. Although no metal complexes of L-type ligands with a 1:2 metal-to-ligand ratio have been previously structurally characterized, a handful of structurally characterized complexes of the physiologically essential divalent metal ions with N2(SR2)4 coordination spheres are available. The metal complexes of type M[N2(SR2)4]2+ [M ) Cu(II),37-39 Ni(II),40-45 Co(II),46 Co(III),43 and Fe(II)47,48] with noncoordinating counterions all exhibit a distorted meridional C2V or trans facial D2h coordination geometry. Table 5 indicates the range of M-N and M-S bond lengths observed in these complexes. The Cu-Naliphatic bond has the largest range of 2.01(1)-2.19(2) Å, and all other divalent first-row transition metal bond distances M(II)-N fall into this range. As might be expected, the average for M-N bond distances increases as one moves down the group 12 family (Zn-N < Cd-N < Hg-N), consistent with the increased metal ion radius. Inspection of Table 5 also indicates that the range of M(II)-S bond distances follows the nearly periodic trend Fe < Ni < Co < Cu < Zn for the first-row transition metal ions. The group 12 metal ions present the feature Zn-S < Cd-S ≈ Hg-S, which was also observed in the six-coordinate group 12 metal crown thioether complexes [M(10S3)2](ClO4)2‚2CH3NO3 [10S3 ) bis(1,4,7-trithiacyclodecane].36 Crystal Structure of Hg(L1)Cl2 (4). As shown in Figure 5, complex 4 has a nearly square-pyramidal Hg(II) coordination geometry (τ ) 0.0649). The square plane is defined by the S(1), S(2), Cl(1), and Cl(2) donor atoms. The sulfur atoms are 0.861 Å from the mean least-squares plane, and the chloride atoms are positioned, on average, a comparable (37) Blake, A. J.; Danks, J. P.; Fallis, I. A.; Harrison, A.; Li, W.-S.; Parsons, S.; Ross, S. A.; Whittaker, G.; Schroder, M. J. Chem. Soc., Dalton Trans. 1998, 3969. (38) Atkinson, N.; Blake, A. J.; Drew, M. G. B.; Forsyth, G.; Gould, R. O.; Lavery, A. J.; Reid, G.; Schroeder, M. J. Chem. Soc., Dalton Trans. 1992, 2993. (39) Atkinson, N.; Blake, A. J.; Drew, M. G. B.; Forsyth, G.; Lavery, A. J.; Reid, G.; Schroder, M. J. Chem. Soc., Chem. Commun. 1989, 984. (40) De Vries, N.; Reedijk, J. Inorg. Chem. 1991, 30, 3700. (41) McAuley, A.; Subramanian, S. Inorg. Chem. 1990, 29, 2830. (42) Chandrasekhar, S.; McAuley, A. Inorg. Chem. 1992, 31, 2234. (43) Blake, A. J.; Reid, G.; Schroeder, M. J. Chem. Soc., Dalton Trans. 1994, 3291. (44) Chiou, S.-J.; Ge, P.; Riordan, C. G.; Liable-Sands, L. M.; Rheingold, A. L. Chem. Commun. 1999, 159. (45) Adhikary, B.; Liu, S.; Lucas, C. R. Inorg. Chem. 1993, 32, 5957. (46) Adhikary, B.; Liu, S.; Lucas, C. R. Inorg. Chim. Acta 1997, 261, 15. (47) Grillo, V. A.; Gahan, L. R.; Hanson, G. R.; Stranger, R.; Hambley, T. W.; Murray, K. S.; Moubaraki, B.; Cashion, J. D. J. Chem. Soc., Dalton Trans. 1998, 2341. (48) Atkinson, N.; Lavery, A. J.; Blake, A. J.; Reid, G.; Schroeder, M. Polyhedron 1990, 9, 2641. (49) Addison, A. W.; Rao, T. N.; Reedijk, J.; Van Rijn, J.; Verschoor, G. C. J. Chem. Soc., Dalton Trans. 1984, 1349.

M-S (Å)

ref(s)

2.248(1)-2.2674(15) 2.460(1)-2.486(1) 2.249(1)-2.268(1) 2.389(2)-2.440(2) 2.487(5)-2.578(5) 2.5300(11)-2.6508(14) 2.6819(5)-2.7359(5) 2.6764(17)-2.7538(17)

46, 47 45 42 39-44 36-38 this work this work this work, 25

distance from the opposite side. The Hg(II) ion is displaced 0.03 Å from the mean least-squares plane toward the chlorides. The largest angles between atoms in the basal plane of the complex are 142.7(1)° for S(1)-Hg-S(2) and 139.3(2)° for Cl(1)-Hg-Cl(2). Five-coordinate Hg(II) complexes exhibiting similar square-pyramidal structures are known,50,51 as are complexes with more trigonal-bipyramidal character.52-54 With L-type ligands, a distorted squarepyramidal geometry is observed in [Ni(L3)Br2] and [Cu(L2)Cl2], with the nitrogen atom, two sulfur, atoms and one halide atom occupying the four basal positions of the square pyramid.8 However, distorted trigonal-bipyramid geometries have also been observed in the complexes [Cu(L8)Cl2] and [Zn(L2)Br2].13,16 Inspection of Tables 3 and 4 shows that the 2.45(1) Å Hg-N and 2.82(5) Å average Hg-S bond lengths in 4 are essentially identical and only 0.10 Å longer, respectively, than their counterparts in complex 1. The intraligand S-Hg-S and N-Hg-S bond angles of 142.7(1)° and 71.4(3)° are ∼4.5% smaller than those in complex 1. The most distinct difference between the ligand conformations in 4 and 1 is the cis and trans orientations, respectively, of the two methyl groups bonded to the two sulfur atoms. The Hg-Cl distances of 2.383(7) and 2.409(6) Å are comparable to the observed range of 2.310-2.428 Å in five-coordinate complexes.51,52,55,56 Crystal Structure of [Cd(L1)Cl(µ-Cl)]2 (5). Complex 5 exists in the solid state as a dimer asymmetrically bridged by two chloride ligands (Figure 6). Each of the Cd(II) centers adopts a distorted octahedral geometry, with one tridentate L1 ligand, two bridging chlorides, and one terminal chloride atom occupying the six positions. The ligand L1 is bound to Cd(II) in an approximately meridional fashion, with one bridging chloride atom in the same plane and the terminal and the other bridging chloride ligand in the axial positions. It is interesting to note that the two L1 moieties are arranged in different configurations. In the dimer, the two methyl groups of one L1 ligand are found cis about the pyridyl plane, but those of the other ligand are trans, resembling Hg(L1)(50) Basu Baul, T. S.; Lycka, A.; Butcher, R.; Smith, F. E. Polyhedron 2004, 23, 2323. (51) Bebout, D. C.; Bush, J. F., II; Crahan, K. K.; Kastner, M. E.; Parrish, D. A. Inorg. Chem. 1998, 37, 4641. (52) Wang, Q.-H.; Long, D.-L.; Huang, J.-S. Polyhedron 1998, 17, 3665. (53) Bebout, D. C.; Garland, M. M.; Murphy, G. S.; Bowers, E. V.; Abelt, C. J.; Butcher, R. J. J. Chem. Soc., Dalton Trans. 2003, 2578. (54) Bebout, D. C.; Ehmann, D. E.; Trinidad, J. C.; Crahan, K. K.; Kastner, M. E.; Parrish, D. Inorg. Chem. 1997, 36, 4257. (55) Bermejo, E.; Castineiras, A.; Garcia, I.; West, D. X. Polyhedron 2003, 22, 1147. (56) Moghimi, A.; Shokrollahi, A.; Shamsipur, M.; Aghabozorg, H.; Ranjbar, M. J. Mol. Struct. 2004, 701, 49.

Inorganic Chemistry, Vol. 45, No. 2, 2006

577

Lai et al. 1

Cl2 and Zn(L )Cl2, respectively [see the following section for Zn(L1)Cl2]. Moreover, the dimeric structure of complex 5 indicates the greater tendency of Cd(II) toward octahedral coordination relative to Hg(II) and Zn(II) and the flexible ligation capacity of halide ligands.57 A similar dimeric structure is common in CdCl2 complexes, such as [Cd(L2)Cl(µCl)]2,14 Cd2(Py2S)2Cl4,7 and Cd2(L)2Cl4 [L ) 1-(5,6-dimethylbenzimidazolyl)-3-benzimidazolyl-2-oxapropane].57 The average Cd-Clbridge bond distance of 2.628(9) Å and Cd-Cd distance of 3.894 Å are comparable to those in chloride-bridged Cd(II) dimers and polymers.58,59 The average Cd-Clterminal bond distance is 2.527(6) Å, similar to those found in six-coordinate Cl-bridged Cd(II) complexes.10,59 The average Cd-S distance of 2.721(4) Å and Cd-N distance of 2.375(9) Å are close to the corresponding 2.713(6) and 2.3803 Å values observed in complex [Cd(L2)Cl(µ-Cl)]2.14 The intraligand N-Cd-S bond angles range from 73.09(10)° to 76.62(10)°. The axial Clbridge-Cd-Clterminal bond angles of 177.01(4)° and 178.38(4)° are slightly smaller than the ideal for an octahedral geometry. Crystal Structure of Zn(L1)Cl2 (6). Complex 6 contains two similar but crystallographically independent Zn(II) centers (Figure 7) in the crystal structure. Each Zn(II) ion is located in a coordination environment consisting of one tridentate L1 and two chloride ligands with slightly more trigonal-bipyramidal character than square-pyramidal {τ ) 0.525 [Zn(1)], 0.604 [Zn(2)]49}. The Zn, N, Cl(1), and Cl(2) atoms are coplanar, defining the equatorial plane, and the two sulfur atoms occupy axial positions. The structure closely resembles that of Hg(L1)Cl2, except for the trans conformation of the two methyl groups with respect to the pyridine plane and a pseudo-C2 axis along the Zn-N bond. The S-Zn-S bond angles of 161.10(2)° and 159.08(2)° deviate significantly from linearity. However, the angles between the atoms in the equatorial plane are approximately ideal, ranging from 113.15(4)° to 129.62(2)°. The average Zn-S bond distance of 2.60(3) Å and Zn-N bond length of 2.081(3) Å are typical for five-coordinate Zn(II) complexes60,61 and comparable to Zn-S(Cys) and Zn-N(His) distances, respectively, in proteins.31-35,62-64 Teixidor et al.16 previously reported the structurally similar complex [Zn(L2)Br2], which has similar Zn-S and Zn-N bond distances. The R groups in complex [Zn(L2)Br2] were found to be trans to each other, as observed in 6. The intraligand S-Zn-S bond angle of 157.0(1)° and N-Zn-S angle of 78.5(7)° also compare well with those in 6. (57) Matthews, C. J.; Clegg, W.; Heath, S. L.; Martin, N. C.; Hill, M. N. S.; Lockhart, J. C. Inorg. Chem. 1998, 37, 199. (58) Pazderski, L.; Szlyk, E.; Wojtczak, A.; Kozerski, L.; Sitkowski, J.; Kamienski, B. J. Mol. Struct. 2004, 697, 143. (59) Loi, M.; Graf, E.; Hosseini, M. W.; De Cian, A.; Fischer, J. Chem. Commun. 1999, 603. (60) Yang, F.-A.; Chen, J.-H.; Hsieh, H.-Y.; Elango, S.; Hwang, L.-P. Inorg. Chem. 2003, 42, 4603. (61) Das, S.; Hung, C.-H.; Goswami, S. Inorg. Chem. 2003, 42, 5153. (62) Karlin, S.; Zhu, Z.-Y.; Karlin, K. D. Biochemistry 1998, 37, 17726. (63) Garcia-Saez, I.; Hopkins, J.; Papamicael, C.; Franceschini, N.; Amicosante, G.; Rossolini Gian, M.; Galleni, M.; Frere, J.-M.; Dideberg, O. J. Biol. Chem. 2003, 278, 23868. (64) Chantalat, L.; Leroy, D.; Filhol, O.; Nueda, A.; Benitez, M. J.; Chambaz, E. M.; Cochet, C.; Dideberg, O. EMBO J. 1999, 18, 2930.

578 Inorganic Chemistry, Vol. 45, No. 2, 2006

Comparisons of Complexes 4-6 to Related Complexes of Physiologically Essential Metal Ions. The crystal structures of [Cu(L8)Cl2],13 [Cu(L2)Cl2],15 [Cu(DAHT)Cl2] {DAHT ) 3,10-dithia-16-azabicyclo[10.3.1]hexadeca-1(16), 12,14-triene},15 [Ni(L3)Br2],8 and [Ni(L3)Br(µ-Br)]29 have been previously reported. [Cu(L8)Cl2] resembles complex 6 with a distorted trigonal-bipyramidal geometry.13 The Cu-S bond distances range from 2.338(3) to 2.431(1) Å [average 2.366(7) Å],13 which are shorter than the Zn-S distances observed in complex 6. The average Cu-N distance of 2.045(6) Å13 is comparable to the average Zn-N distance of 2.080(7) Å in 6. The average S-Cu-S bond angles of 158.7(8)° is also comparable to the S-Zn-S angle of 160.09(2)° observed in complex 6. The complexes [Cu(L2)Cl2],15 [Cu(DAHT)Cl2],15 and [Ni(L3)Br2]8 exist in the solid state as monomers adopting distorted square-pyramidal coordination geometries much like 6. The complex [Ni(L3)Br2]8 has an average Ni(II)-S distance of 2.383(7) Å and a Ni(II)-N distance of 2.057(5) Å that compares well with the above Cu(II) complexes. However, a cis arrangement of the R groups of the ligand was observed, as in complex 4. The R groups of the ligands in complexes [Cu(L2)Cl2]15 and [Cu(L8)Cl2]13 are both arranged in trans fashion as observed in complex 6. The dimeric complex [Ni(L3)Br(µ-Br)]29 has a distorted octahedral coordination geometry for Ni(II), which is similar to those of 5 and [Cd(L2)Cl(µ-Cl)]2.14 The average Ni-N and Ni-S bond distances are 2.0557(5) and 2.443(7) Å, respectively, and their R groups are both trans to each other. Crystallographic database searches revealed four other, non-L-based Cu(II)65-68 complexes and one Zn(II)69 complex with a N(SR2)2X2 coordination sphere. The Cu-S bond distances in these complexes range from 2.313(2) to 2.369(2) Å with an average of 2.334(9) Å, which falls in the observed range of the above Cu(II)-L complexes. The average CuNpyridyl and Cu-Naliphatic bond distances are 2.025(9) and 2.174(4) Å, respectively. The average Zn-S and Zn-N bond distances are 2.618(1) and 2.080(3) Å, which are comparable to the corresponding 2.686(7) and 2.083(5) Å values observed in the complex [Zn(L2)Br2].16 As discussed above, the halide salts used to study the coordination chemistry of the L-type ligands are limited to CuX2, NiX2 and ZnX2, with only a few cases of heavy metal salts. Also, with only a few exceptions, such ligands are found coordinated to metal ions in a tridentate fashion. The mononuclear complexes M[N(SR)2X2] with tridentate coordination all adopt distorted square-pyramidal or trigonalbipyramidal coordination geometries. Furthermore, the halide counterions are found in the primary coordination sphere, which could prevent a second ligand from binding. (65) Arca, M.; Blake, A. J.; Lippolis, V.; Montesu, D. R.; McMaster, J.; Tei, L.; Schroder, M. Eur. J. Inorg. Chem. 2003, 1232. (66) Ferlay, S.; Jouaiti, A.; Loi, M.; Hosseini, M. W.; De Cian, A.; Turek, P. New J. Chem. 2003, 27, 1801. (67) Agnus, Y.; Louis, R.; Weiss, R. J. Am. Chem. Soc. 1979, 101, 3381. (68) Weber, G. Inorg. Chim. Acta 1982, 58, 27. (69) Nekola, H.; Rehder, D. Inorg. Chim. Acta 2002, 337, 467.

Zn Triad Coordination of 2,6-Bis(methylthiomethyl)pyridine

Figure 8. Methylene (Hf) region of the 1H NMR spectra for L1 at -40 °C in acetonitrile-d3 as a function of added Zn(ClO4)2 (left), Cd(ClO4)2 (middle) and Hg(ClO4)2 (right) with nominal [M2+]/[L1] ratios of 0.125, 0.250, 0.375, and 0.500. The nominal total M2+ concentration was 2 mM.

Solution-State 1H NMR Spectroscopy. Metal ion binding to L1 with the perchlorate and chloride salts of divalent zinc triad metals was studied in the solution state using 1H NMR spectroscopy. The room-temperature acetonitrile-d3 solution 1 H chemical shifts for complexes 1-6 are generally downfield of equivalent resonances for the free ligand. Downfield shifts are often observed upon complexation to metal ions and are attributed to the deshielding influence of σ donation to the metal center. Acetonitrile-d3 solutions of the two series of complexes 1-3 and 4-6 exhibited similar 1H chemical shifts with no apparent periodic trends. Furthermore, the asymmetric ligand binding observed in all six crystal structures was not observed in solution. Each complex had a single resonance for each symmetry-related proton in the free ligand. The small differences between M(II)-S bond lengths ( 1, but the 199Hg coupling satellites (Figure 9b) became exchange-broadened, indicating ligand exchange between solvated Hg2+ and Hg(L1)(NCCH3)x2+. In contrast, the proton NMR spectra of Cd(L1)22+ and Zn(L1)22+ remained essentially invariant in solutions with [M(ClO4)2]/[L1] > 0.625. This suggests that M(L1)22+ is more thermodynamically stable than M(L1)(NCCH3)x2+ for M ) Cd2+ and Zn2+ over the range of [M(ClO4)2]/[L1] ratios examined (Figure S4). However, the changes in J(199Hg1H) are more definitive than the changes in proton chemical shift for the interconversion of Hg(L1)22+ and Hg(L1)(NCCH3)x2+. Because J(111/113Cd1H) was never detected for the Cd2+ complexes and zinc does not have a favorable isotope for NMR detection, the M/L1 stoichiometry of the Cd2+ and Zn2+ complexes at [M(ClO4)2]/[L1] > 0.5 is ill-defined in solution. Comparable solution equilibria have been reported for the Hg(II) coordination chemistry of related tridentate ligands.7,21 The observed 3J(199Hg1H) values for Hg(L1)(NCCH3)x2+ of 105 and 95 Hz are among the largest coupling constants reported for thioether-ligand-containing Hg(II) complexes.7,71 Reports of longer-range J(199Hg1H) couplings in Hg(II) coordination compounds are rare. Those 4-5J(199Hg1H) couplings reported here fall in the known range of 8-30 Hz for coupling between 199Hg and 1H separated by four or five bonds.7,21,53,72-74 (70) Bebout, D. C.; Stokes, S. W.; Butcher, R. J. Inorg. Chem. 1999, 38, 1126. (71) McCrindle, R.; Ferguson, G.; McAlees, A. J.; Parvez, M.; Ruhl, B. L.; Stephenson, D. K.; Wieckowski, T. J. Chem. Soc., Dalton Trans. 1986, 2351. (72) Utschig, L. M.; Bryson, J. W.; O’Halloran, T. V. Science 1995, 268, 380. (73) Bebout, D. C.; Bush, J. F., II; Crahan, K. K.; Bowers, E. V.; Butcher, R. J. Inorg. Chem. 2002, 41, 2529.

580 Inorganic Chemistry, Vol. 45, No. 2, 2006

L1

The coordination chemistry of binding to divalent group 12 metal chlorides was also investigated by 1H NMR spectroscopy. Although 5 had a different structure than 4 and 6 in the solid state, the 1H NMR spectra of the three complexes were strikingly similar. The observed proton chemical-shift differences for 4 and 5 in acetonitrile-d3 were less than 0.02 ppm, and resonances for 6 were within 0.22 ppm of those for the other chloride complexes. The chemicalshift differences between these complexes when recorded in DMSO-d6 solutions were even smaller (∆δ e 0.08 ppm, data not shown). These data suggest that the complexes are structurally similar in solution. In this context, dimeric and monomeric structures are likely extremes in a continuum that reflect subtle differences in packing constraints in the solid state or solubility differences. Furthermore, no J(M1H) couplings were found for complexes 4 and 5 at accessible temperatures with [M2+] e 2 mM in either solvent. Previous work suggests that intermolecular ligand exchange of d10 metal ions with coordinating counterions such as chloride ions occurs particularly rapidly because of reductions in the available charge density. Precedent for the detection of 199 Hg-1H couplings in coordination compounds of HgCl2 is limited to tetradentate chelating ligands.51,53,54 Conclusion Solid-state characterization of six divalent group 12 metal complexes of L1 revealed tridentate binding modes with comparable ligand geometries except for the relative orientation of the thioether methyl groups. The binding of two ligands to the same metal center was first observed for the pyridine-based dithioether ligands in complexes M(L1)2(ClO4)2 (M ) Hg, Cd, Zn). The three complexes all display distorted octahedral structures in the solid state with the two ligands in similar conformations. Complexes Hg(L1)Cl2 and Zn(L1)Cl2 exist in the solid state as monomers with distorted square-pyramidal and trigonal-bipyramidal geometries, respectively. [Cd(L1)Cl(µ-Cl)]2 exists in the solid state as a dimer of Cd(II) centers with a distorted octahedral coordination geometry. The slight ligand binding asymmetry observed in the solid state for these six complexes was not observed in solution because of the ease of rearranging the coordination sphere of d10 metal ions in the absence of crystal restraints and the low energy barrier to thioether inversion. Furthermore, the differences between the proton NMR spectra of 4-6 were negligible, suggesting that the different structures in the solid state were artifacts associated with crystallization. There were, however, interesting differences between the solutionstate coordination chemistries of zinc triad perchlorate complexes of L1. The relative rates of intermolecular ligand exchange for ions of type ML22+ decreased in the order Cd(II) > Zn(II) > Hg(II). Both 1:1 and 1:2 complexes of Hg(II) with L1 were in slow exchange on the couplingconstant time scale below -20 °C, but slow exchange conditions could not be found for the equivalent Cd(II) complexes. (74) McWhinnie, W. R.; Monsef-Mirzai, Z.; Perry, M. C.; Shaikh, N.; Hamor, T. A. Polyhedron 1993, 12, 1193.

Zn Triad Coordination of 2,6-Bis(methylthiomethyl)pyridine

Despite intense study, the detailed molecular mechanisms for the toxicology of heavy metals remain largely unresolved. Although it is generally appreciated that Hg(II) has different toxicological effects than Cd(II), the solution-state research described here offers the fundamental insight that slow intermolecular exchange of Hg(II) between multidentate coordination sites encountered relative to Cd(II) might contribute to the differences in their toxicological profiles. Similarly, physiologically essential Zn(II) is bound more transiently to the ligand investigated, suggesting that Zn(II) can be more readily displaced when bound to inappropriate sites. Comparisons of the solution-state behavior of complexes 1-3 suggest that Hg(II) might compete more effectively for the native metal-binding sites of some proteins than Cd(II) or even Zn(II). Furthermore, slow intermolecular ligand exchange for the lower-coordination-number 1:1 Hg2+/ L1 complex suggests that Hg(II) might not require preformed metal-binding sites with high coordination numbers to have toxic effects. Adventitious coordination of Hg(II) to arrangements of metal-binding amino acids on the surface of a protein might allosterically interfere with native functions or misdirect the folding of specific proteins. Differences in Hg(II), Cd(II), and Zn(II) binding to native metal-binding

sites and adventitious sites might contribute to their vastly different physiological impacts. Further examination of group 12 metal coordination to multidentate ligands should provide further insight regarding the differences between their bioactivities. Acknowledgment. This work was supported in part by the Petroleum Research Fund under Grant 40151-B3 and the National Science Foundation under Grant 0315934. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation. We also thank the Camille and Henry Dreyfus Foundation for a Scholar Fellow award to D.C.B. for support of S.M.B. and the research. R.J.B. acknowledges the DoDONR instrumentation program for funds to upgrade the diffractometer and the NIH-MBRS program for funds to maintain the diffractometer. Supporting Information Available: X-ray crystallographic data (CIF), additional NMR data. This material is available free of charge via the Internet at http://pubs.acs.org. IC051091+

Inorganic Chemistry, Vol. 45, No. 2, 2006

581